Integrative Biology

View Article Online View Journal

Accepted Manuscript

This article can be cited before page numbers have been issued, to do this please use: J. J. Linderman, N. A. Cilfone, E. Pienaar, C. Gong and D. E. Kirschner, Integr. Biol., 2015, DOI: 10.1039/C4IB00295D.

This is an Accepted Manuscript, which has been through the Royal Society of Chemistry peer review process and has been accepted for publication. Accepted Manuscripts are published online shortly after acceptance, before technical editing, formatting and proof reading. Using this free service, authors can make their results available to the community, in citable form, before we publish the edited article. We will replace this Accepted Manuscript with the edited and formatted Advance Article as soon as it is available. You can find more information about Accepted Manuscripts in the Information for Authors. Please note that technical editing may introduce minor changes to the text and/or graphics, which may alter content. The journal’s standard Terms & Conditions and the Ethical guidelines still apply. In no event shall the Royal Society of Chemistry be held responsible for any errors or omissions in this Accepted Manuscript or any consequences arising from the use of any information it contains.

www.rsc.org/ibiology

Page 1 of 37

Integrative Biology View Article Online

DOI: 10.1039/C4IB00295D

3   4   5   6   7   8   9   10   11   12   13   14   15   16   17   18   19   20   21   22   23   24   25   26   27   28   29   30   31   32   33   34   35   36   37   38   39   40   41   42   43   44   45   46  

A Multi-scale Approach to Designing Therapeutics for Tuberculosis Jennifer J. Linderman#, Nicholas A. Cilfone#, Elsje Pienaar#*, Chang Gong&, Denise E. Kirschner* #

Dept. Chemical Engineering, Univ. of Michigan. *Dept. Microbiology and Immunology, Univ. Michigan Medical School. &Dept. Computational Medicine and Bioinformatics, Univ. Michigan Medical School.

Abstract Approximately one third of the world’s population is infected with Mycobacterium tuberculosis. Limited information about how the immune system fights M. tuberculosis and what constitutes protection from the bacteria impact our ability to develop effective therapies for tuberculosis. We present an in vivo systems biology approach that integrates data from multiple model systems and over multiple length and time scales into a comprehensive multi-scale and multi-compartment view of the in vivo immune response to M. tuberculosis. We describe computational models that can be used to study (a) immunomodulation with the cytokines tumor necrosis factor and interleukin 10, (b) oral and inhaled antibiotics, and (c) the effect of vaccination. Introduction Tuberculosis (TB), a deadly infectious disease caused by the bacterium Mycobacterium tuberculosis, results in 1-2 million deaths per year worldwide. In 2013, an estimated 9 million new cases were diagnosed[1]. Control of the TB epidemic is limited by a complex and prolonged antibiotic regimen, development of antibiotic resistance, the lack of an effective vaccine and, more generally, by our incomplete understanding of the hostpathogen dynamics that underlie the disease, its progression, and treatment[2, 3]. Many of the challenges to the development of therapies for TB are captured by the following as-yet-unanswered questions: (1) What is the immune response to M. tuberculosis infection, and why does it often fail to eliminate the infection? Figure 1 shows key aspects of the immune response to M. tuberculosis. M. tuberculosis is a respiratory pathogen that primarily causes infection in the lungs in adult humans and is transmitted via aerosolized droplets of bacteria from an infectious individual. Upon inhalation, bacteria reach the pulmonary alveoli and are phagocytosed by macrophages that line the alveolar space. Although some bacteria may be destroyed by macrophages through antimicrobial mechanisms, M. tuberculosis has evolved ways to evade protective host immune

 

1  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

1   2  

Integrative Biology

Page 2 of 37 View Article Online

47   48   49   50   51   52   53   54   55   56   57   58   59   60   61   62   63   64   65   66   67   68   69   70   71   72   73   74   75   76   77   78   79   80   81   82   83   84   85   86  

mechanisms (e.g. by preventing phagosome fusion with the lysosome), and as a consequence is able to multiply within macrophages[4, 5]. Dendritic cells (DCs), another phagocytic cell type, also internalize M. tuberculosis. DCs migrate through lymphatics to lung-draining lymph nodes (LNs) to prime an adaptive immune response. About 2-4 weeks after the initial infection, effector T cells, attracted by chemokines and pro-inflammatory cytokines from the lung site of infection, migrate back to lungs via the blood to mount an M. tuberculosis-specific immune response. The net result of these events is the formation of granulomas, roughly spherical collections of immune and lung cells, bacteria and infected cells.

87   88   89   90   91  

(2) Why do some individuals develop latent disease while others develop active disease? Humans and non-human primates infected with M. tuberculosis have multiple granulomas, from a few to ~25 granulomas[10, 12, 17]. The manifestation of the disease in an individual depends on how well the collection of granulomas can control infection. Following an initial infection with M. tuberculosis, ~10% of

In TB, the battle between host and microbe plays out at the level of the granuloma. A classic caseous granuloma consists of a central necrotic area surrounded by layers of macrophages and then a smaller cuff of lymphocytes[4, 6]. The lymphocytic cuff primarily contains both CD4+ and CD8+ T cells, but other cell types, including B cells, neutrophils, DCs and fibroblasts are also observed[7, 8]. There are also many molecular mediators of granuloma dynamics, including cytokines interferon-γ (IFNγ), tumor necrosis factor-α (TNF), and interleukin-10 (IL-10) and chemokines CXCL9/10/11, CCL2 and CCL5. Some, like IFN-γ, have been shown to be necessary to M. tuberculosis infection control, while others remain controversial. None have been shown to be sufficient for infection control. A central feature of almost all granulomas is a caseous necrotic center (dead immune cells and lung tissue), often trapping large numbers of bacteria that are unable to grow due to hypoxic conditions. The role of a granuloma from a host-centric point of view is to contain infection, destroy bacilli, and limit pathology. From the bacterial point of view, however, the granuloma may serve as a niche for survival. If all granulomas present are capable of inhibiting or killing most mycobacteria present, humans develop a clinically latent infection. However, if a granuloma does not control bacterial growth, infection progresses, granulomas enlarge, and bacteria seed new granulomas; this results in progressive pathology and disease, i.e. active TB[5, 9-13]. Mechanisms that lead to an inability of the immune response to completely eliminate the pathogen are unknown but appear to be both host- and bacteria- related, making it difficult to identify those that would be suitable to manipulate for therapeutic purposes. Further, the immune response is necessarily limited; an overly-enthusiastic immune response, while possibly eliminating the bacteria, can do considerable damage to host lungs[11, 12, 14-16]. Perhaps latent disease is simply a compromise that, for the most part, works. However, a third of the world’s population is thought to have latent TB, providing a huge reservoir of contagion (contributing to the pool of active disease through reactivation); treating latent TB will be essential to the ultimate eradication of a disease that claims millions of lives each year[1].

 

2  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Page 3 of 37

Integrative Biology View Article Online

92   93   94   95   96   97   98   99   100   101   102  

humans develop primary (active) TB, ~90% develop latent infection, and a few individuals likely clear the disease. Reactivation TB refers to the situation in which an individual with latent TB later develops active disease either due to reactivation of existing infection or a reinfection event[18]; there is a 10% per lifetime risk that can be greatly increased with immune-compromising events. It may be that some individuals with latent TB will never, or only rarely, develop reactivation TB, while for others the risk is much greater, i.e. there may be a spectrum of latency[19]. The factors that control these different outcomes are not well-understood. We do know that interfering with the immune system, either pharmacologically by delivering antiTNF therapies (used in the treatment of some autoimmune diseases) or pathologically in the case of HIV-1 co-infection, both increase the risk of reactivation[5, 20-24].

103   104   105   106   107   108   109   110   111   112   113   114   115   116   117   118   119   120   121   122   123   124   125   126   127   128   129   130   131   132   133   134   135   136   137  

(3) Why is a long time course of antibiotics needed, and why do antibiotics often fail? Standard therapy for active TB includes an initial combination of 3-4 first-line oral antibiotics for two months followed by another 4-7 months of 2 oral antibiotics[25]. Long treatment periods appear to be required because of the presence of phenotypically drug-tolerant ‘persister’ bacteria, slow bacterial growth rates and high bacterial loads[26-29]. Known obstacles to treatment success (including patient non-compliance, drug toxicity, relapse, and drug resistance) are thought to be, at least in part, a result of the unusually long treatment regimens[1, 2, 30, 31]. The complex nature of the site of infection, namely granulomas, presumably further complicates treatment, as the dense and heterogeneous tissue itself may present an obstacle for antibiotics to reach the site of infection. Worldwide, TB has an 87% treatment success rate in new cases, leaving more than 1 million new patients without cure in 2012 1. Thus there is a great need for both shorter treatment regimens and new antibiotics[32-34]. (4) Why is there no vaccine against TB? To date there is still no efficacious vaccine against M. tuberculosis, although ~30 vaccines are in various stages of testing and clinical trials (www.aeras.org/annualreport)[35-39]. These trials are expensive, difficult, and time consuming to perform, and many result in a null outcome. The development of effective vaccines against TB is challenging as the immune responses necessary for prevention of infection are still unknown. Although many infants are vaccinated at birth with BCG (an attenuated M. bovis), this does not prevent infection or development of TB after childhood. Data suggest an effective vaccine must generate memory cells to multiple M. tuberculosis antigens that are expressed at multiple stages during infection[40]. However, there are currently no comprehensive approaches or tools that could significantly advance the development of vaccines. (5) What other approaches for treatment of disease might be explored? Approaches that would augment antibiotic therapy following infection with M. tuberculosis are now under consideration. Combining immune modulation (“immunomodulation”) with antibiotics is a potential strategy for enhancing treatment of TB[41-43]. Immunomodulation here refers to the addition/subtraction of cells and/or molecules (e.g. cytokines) to enhance the immune response. It stands to reason that boosting the immune system while reducing bacterial load could lead to more rapid control of infection. Several strategies have been tried in murine models (IFN-γ,  

3  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Integrative Biology

Page 4 of 37 View Article Online

138   139   140   141   142   143   144   145   146   147   148   149   150   151   152   153   154   155   156   157   158   159   160   161   162   163   164   165   166   167   168   169   170   171   172   173   174   175   176   177   178   179   180   181   182   183  

GM-CSF, TNF, IL-12)[41, 42] and a few in humans (IL-2, GM-CSF, TNF, IFN-γ)[41, 43], but results are inconclusive. Again, the complexity of the immune response makes it difficult to identify which mechanisms are appropriate to modulate to increase control of infection while simultaneously minimizing tissue damage and extensive inflammation. Appropriate delivery to granulomas and proper timing, drug combinations and dosing are all likely to be key factors in successful therapy. Underlying these five questions is a common theme: we currently have only limited insight into how the immune system fights M. tuberculosis and what constitutes protection from the bacteria. As a result, it is difficult to know how best to develop treatments and to approach vaccine development for TB. There are many reasons why these questions have been and remain difficult to address. Two reasons are particularly relevant for the discussion in this issue on in vivo systems biology. First, it should be clear from the above that, at a minimum, the lungs, draining LNs, blood, and lymphatic system participate in the host-pathogen dynamics that describe M. tuberculosis infection and its treatment (Figure 1), so it is difficult to study the disease “in a dish”. Most experimental studies focus on a single biological (length and/or time) scale of interest, e.g. examination of immune cells in the blood or a particular signaling pathway. Figure 2 highlights the different spatiotemporal scales at which host-pathogen dynamics operate. The smallest spatial scale shown, the molecular scale, also represents the fastest time scale. Receptor/ligand binding and trafficking as well as signal transduction pathways are included at this scale. Examples of assays that generate data for this scale include flow cytometry for receptor expression and fluorescently tagged reporters for gene expression[44-47]. The actions of individual cells, e.g. apoptosis, movement or secretion, are tracked at the cellular scale. Experiments such as microfluidic chemotaxis assays, TUNEL staining, and ELISA assays measuring cytokine production generate data for this scale[48-51]. The major event occurring at the tissue scale in TB is the formation of granulomas; necrosis and fibrosis are also tissue-level outcomes in TB. Experiments at this scale examine gross-pathology, histology, bacterial loads, cellular distribution and fibrosis[7, 12, 17, 52]. These tissue-scale events evolve over periods of weeks to years in humans. The largest spatial scale shown here is the organ scale. Here cells can traffic between the site of infection through blood to draining LNs and back again as well as to other body sites. Thus to understand the complex in vivo immune response to M. tuberculosis, it will be important to integrate information from experiments performed at multiple scales and on multiple physiological compartments (lung, blood, lymphatics, LNs). Second, the “design space” of both potential experiments and potential therapies is enormous. For example, the identities, dosing schedules, and concentrations of multiple antibiotics, cytokines and/or other immunomodulators can be varied across a wide range[53]. Animal models have been used for nearly 100 years in the study of TB and have provided much useful data. However, the animal models that are easiest to work with may not fully capture human disease. Mice are most commonly used because of the availability of reagents, genetically modified animals, and ease of use. But there is no true latent infection in mice; they become chronically and progressively infected, and eventually succumb to the disease. In addition, mouse granulomas are substantially

 

4  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Page 5 of 37

Integrative Biology View Article Online

184   185   186   187   188   189   190   191   192   193   194   195   196   197   198   199   200   201   202   203   204   205   206   207   208   209   210   211   212   213   214  

different from human granulomas in terms of structure and organization[5, 6]. Other small organisms, e.g. guinea pigs, rabbits and zebrafish, have their own advantages and disadvantages[6, 54-56]. Recently, non-human primates, in particular Cynomolgus macaques, have emerged as the animal model most similar to humans in terms of spectrum of disease outcomes and pathology[6, 12]. The cost, technical, and ethical issues of working with macaques means that the number of animals studied, i.e. the fraction of design space that can be explored, is necessarily small. It is even more difficult and expensive to evaluate new therapies or vaccines in human clinical trials. Thus there is a crucial need for an approach that can efficiently narrow the design space of potential experiments and be used to identify, test, and optimize new therapies for TB.

215   216   217   218   219   220   221   222   223   224   225   226   227  

As shown in Figure 3, there are three central elements of our approach for following granuloma formation and function[63]. First, we use an agent-based model (ABM) to describe cellular behavior, including recruitment to the lung, changes of state (activation, infection, etc.), and movement (Figure 3A). Cells (agents) included are macrophages and T cells which can have multiple states (e.g. infected, activated, etc.). Bacteria are not represented as agents but rather as continuous functions in the extra- or intra-cellular environment. We track three different bacterial populations in our model: intracellular replicating, extracellular replicating and extracellular non-replicating bacteria. The simulation environment is two-dimensional and represents a 4-16 mm2 cross-section of lung tissue. Probabilistic interactions between immune cells and with bacterial populations are described by a well-defined set of rules between immune cells and M. tuberculosis in the lung. Each simulation follows events over several hundred days, building over time to track thousands of individual cells (agents).

We are convinced that a systems biology approach that integrates data from multiple model systems and over multiple length and time scales into a comprehensive multi-scale and multi-compartment view of the in vivo immune response to M. tuberculosis is necessary. This approach will allow us to identify and understand the mechanisms underlying host-pathogen dynamics and to improve on and identify new therapies. In this paper, we discuss our computational models of M. tuberculosis infection, focusing on the multi-scale and multi-compartment influences that lead to granuloma formation and influence granuloma function – the ability to contain infection, and in the presence of minimal tissue damage and inflammation. We believe computational models provide valuable tools (among others) to aid in addressing the questions introduced above. Computational models of granuloma formation and function As explained above, and in Figure 2, the “readout” of the lower and higher scale events – signaling pathways, cellular actions, and cellular input from lymph nodes – are granulomas (occurring at the tissue scale) that may contain infection and may be accompanied by significant tissue damage and inflammation. Thus we have developed computational models of granuloma formation and function that are formulated such that information can be continually exchanged across scales and in both (higher/lower scales) directions[14, 20, 57-62].

 

5  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Integrative Biology

Page 6 of 37 View Article Online

228   229   230   231   232   233   234   235   236   237   238   239   240   241   242   243   244   245   246   247   248   249   250   251   252   253   254   255   256   257   258   259   260   261   262   263   264   265   266   267   268   269   270   271   272   273  

Second, we capture receptor/ligand binding and trafficking and intracellular signaling events with ordinary differential equations (ODEs) that are solved within each agent (Figure 3B)[14, 57, 58, 61, 63]. For instance, the model can capture receptor-ligand binding and trafficking of cytokines, such as tumor necrosis factor-α (TNF) or interleukin-10 (IL-10), using ODEs[14]. The detail required in the model at this scale is determined by the questions being asked. For example, a detailed description of cytokines is necessary when trying to understand how cytokine availability and signaling contribute to infection control. If focus shifts to elucidating the dynamics of antibiotic treatment in granulomas, a detailed description of cytokines may not be necessary. Thus we use an approach we term tunable resolution, formulating fine-grained (detailed) and coarsegrained (less detailed) descriptions of the biological events occurring and toggling between these levels of resolution as needed.[64, 65] Third, we describe the diffusion of particular chemokines, cytokines, and other soluble ligands (e.g. anti-TNF antibodies, antibiotics) by solving the relevant partial differential equations (Figure 3C). Equations and parameters for these portions of the model are based on extensive biological data. The three model elements are linked, allowing information to be continually exchanged across scales (Figure 3)[63]. Thus, our overall computational model of M. tuberculosis infection and granuloma formation is hybrid (formed from different mathematical formalisms). It is also multi-scale, incorporating molecular and cellular events explicitly with tissue-scale behavior (granuloma formation) as an emergent feature of the model (Figure 4). Among other tools, we use uncertainty and sensitivity analyses techniques to understand the relative importance of particular processes to granuloma formation and function[66]. Although not discussed in detail here, our granuloma models have been developed, calibrated, and validated using extensive data from mice and primates. Figure 4A shows an example of model calibration to the number of bacteria (CFU, or colony-formingunits) per granuloma in non-human primates[10, 17, 61]. Figures 4B and 4C show snapshots from two simulations using different but physiologically realistic parameter values. We can predict features that map to a wide spectrum of those observed in primates, including granulomas that are able to contain bacteria (Figure 4B), granulomas that show bacterial overgrowth and dissemination (Figure 4C), and granulomas that clear bacteria completely, sometimes with extensive inflammation (not shown). Investigating therapeutic approaches Our computational models of granuloma formation and function can be used to probe interventions that improve the ability of a granuloma to contain, or even eliminate, bacteria while minimizing tissue damage and inflammation. Several potential interventions with actions at different scales are shown in Figure 2 (interventions 1-6) and are discussed below. In each case, interventions at one location, or in one type of molecule or cell, impact events at other length and time scales, and we are especially interested in their effect on our main “readout”, granuloma formation and function. Below we describe four examples to highlight ways in which our approach can help the discovery of therapeutic interventions for M. tuberculosis infection. They are organized

 

6  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Page 7 of 37

Integrative Biology View Article Online

274   275   276   277   278   279   280   281   282   283   284   285   286   287   288   289   290   291   292   293   294   295   296   297   298   299   300   301   302   303   304   305   306   307   308   309   310   311   312   313   314   315   316   317   318   319  

according to their scale to emphasize that there are multiple levels that should be considered when designing interventions opening up new avenues for exploration. 1. Immunomodulation focused on IL-10 and TNF Data from human, animal, and mathematical models have demonstrated that proinflammatory cytokines, such as TNF, are essential to an efficient antimicrobial response against M. tuberculosis infection[22, 67, 68]. However, many anti-inflammatory cytokines are also present in granulomas[4, 5]. In particular, the regulatory cytokine IL10 is of interest since it functions to inhibit cytokine and chemokine production (specifically TNF)[69-72]. It has recently been proposed that a balance of pro- and antiinflammatory mediators (such as TNF and IL-10) in granulomas is an essential component of an efficient antimicrobial response with limited host-induced tissue damage[5, 73, 74]. Understanding how cytokines contribute to infection control at a single granuloma level has been difficult due to the myriad of cellular sources, differences among animal models, and limitations of detection methods for these mediators. From a therapeutic standpoint, we simply do not know whether manipulation of pro- and anti-inflammatory cytokines in granulomas would be useful in affecting infection outcomes. In order to explore the therapeutic value of modulating cytokine levels, i.e. immunomodulation, we developed a version of our multi-scale computational model that incorporates IL-10 and TNF cytokine dynamics across multiple temporal and spatial scales (Figure 3)[14, 57, 61]. This model describes cytokine secretion, diffusion, degradation, and receptor-ligand binding and trafficking. We link these mechanisms across scales by allowing dynamics within each scale to influence behavior at other scales. For example, TNF binding and subsequent internalization affect TNF concentrations in the granuloma environment, affecting cellular apoptosis/necrosis. This systems biology-based approach allows us to explore the effect of immunomodulation strategies at the individual granuloma scale by performing virtual IL-10 knockouts, temporal IL-10 knockouts, and perturbing the balance of TNF and IL-10 levels in granulomas. We first performed a virtual IL-10 deletion (referred to as IL-10 K/O) at the initialization of infection by setting IL-10 synthesis rates for all cells (agents) to zero (Figure 2, intervention 4). We observed a significant change (~2-fold increase) in the number of granulomas that achieve sterility (granulomas that kill all bacteria within) in the IL-10 K/O simulations as compared to the wild-type (WT) simulations (Figure 5A). The mean bacterial load per granuloma at 200 days post-infection is reduced ~1.75-fold in IL-10 KO simulations. However, when sterile granulomas are removed from the analysis of mean bacterial loads per granuloma, there is no significant difference between IL-10 KO simulations and WT simulations (Figure 5A). Thus, the model predicts that reduced bacterial loads in IL-10 KO simulations are due solely to the increased number of granulomas that are successfully able to sterilize bacteria. This suggests that IL-10 is a key regulator of granuloma sterility and that IL-10 focused treatment strategies might be able to improve infection outcome.

 

7  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Integrative Biology

Page 8 of 37 View Article Online

320   321   322   323   324   325   326   327   328   329   330   331   332   333   334   335   336   337   338   339   340   341   342   343   344   345   346   347   348   349   350   351   352   353   354   355   356   357   358   359   360   361   362   363   364   365  

In order to better understand whether IL-10 could be an effective therapeutic strategy, we performed virtual IL-10 deletions at days 25, 50, 75, and 100 post-infection. We observed an initial increase in the number of sterile granulomas depending on when IL10 was removed from the system (Figure 5B), indicating that an increase in sterile granulomas due to deletion of IL-10 is a phenomenon that primarily occurs during the early immune response to M. tuberculosis. Unfortunately, the early increase in granuloma sterilization is coupled with increases in inflammation and tissue damage (not shown). Taken together, these predictions suggest that any therapeutic value of modulating IL-10 levels in granulomas may be present only at early times post-infection. However, because most patients typically present symptoms weeks to months after initially becoming infected with M. tuberculosis, this strategy is unlikely to be implemented in a clinical setting. It does, however, point to the importance of unbridling the immune response early in TB, which might be accomplished via a vaccine. Similarly, modulating levels of TNF in granulomas could prove useful as a therapeutic strategy. In work with earlier generation granuloma models, we showed that altering TNF levels, TNF receptor internalization capabilities, or rates in the TNF-induced NFkB signaling pathway could alter granuloma outcomes, e.g. containment vs. dissemination of bacteria (Figure 2 – interventions 1-3)[20, 21, 57, 58]. Next, we modulated cellular production rates of both IL-10 and TNF within our granuloma simulations, thus changing the balance of TNF and IL-10 during infection [14, 61]. When the ratio of TNF to IL-10 in a granuloma is less than ~0.1, anti-inflammatory mechanisms dominate the immune response. We observe elevated bacterial loads (Figure 5C) with no granulomas achieving sterilization of bacteria. At the same time, however, the presence of caseation at the granuloma’s center , a measure of tissue damage, was reduced nearly 10-fold (Figure 5D). Conversely, when the ratio of TNF to IL-10 is greater than ~1.0, the immune response to M. tuberculosis infection is dominated by the pro-inflammatory response. In this case, significantly more granulomas are able to successfully sterilize (Figure 5C), but increased antimicrobial activity comes at the cost of increased tissue damage (Figure 5D). If the ratio of TNF to IL-10 is between these two extremes a trade-off exists between granuloma sterilization and tissue damage[14]. Thus controlling the balance of TNF and IL-10 in a granuloma, using exogenous antibodies or cytokines, could be an effective therapeutic strategy to shift granuloma outcomes from bacterial containment to sterilization. However, modulation of TNF and IL-10 must be done in a precise and perhaps a time-limited way to limit excessive inflammation and tissue damage. These results suggest that immunomodulation strategies focusing on balancing pro- and antiinflammatory cytokines such as TNF and IL-10 could have significant therapeutic value, perhaps in combination with antibiotics. 2. Antibiotics Our computational approach can also be used to examine the action of oral antibiotics during M. tuberculosis infection (Figure 2 – intervention 6). In particular, we wanted to understand the failure of current antibiotic treatments and to provide a tool for assessing how antibiotics or antibiotic dosing regimens might be altered to improve efficacy. To do

 

8  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Page 9 of 37

Integrative Biology View Article Online

366   367   368   369   370   371   372   373   374   375   376   377   378   379   380   381   382   383   384   385   386   387   388   389   390   391   392   393   394   395   396   397   398   399   400   401   402   403   404   405   406   407   408   409   410  

this, we took a systems pharmacology approach, incorporating pharmacokinetic (PK) and pharmacodynamics (PD) elements into our computational models of granuloma formation and function (Figure 6)[75]. Current first-line antibiotics for TB are isoniazid (INH), rifampin (RIF), pyrazinamide (PZA) and ethambutol (EMB). Because INH and RIF are typically administered during the entire regimen and are arguably the most well-studied of the group[25], we focus on them here. We incorporated PK and PD models of orally-dosed INH and RIF into our computational model of granuloma formation[75]. PK are described by a classical two compartment (plasma and body) model with two absorption compartments (Figure 6B)[76]. Antibiotic concentrations in the plasma compartment of the PK model are used to determine the movement of antibiotics into or out of the granuloma simulation grid (Figure 6A). Vascular permeation of antibiotics onto the ABM occurs at grid compartments designated as vascular sources (Figure 3A), and depends on antibiotic concentration gradients between blood and vascular source grid compartments. Antibiotics can diffuse and degrade on the simulation grid and enter host cells – which we refer to as granuloma PK. PD are modeled using a Hill curve [77] and antimicrobial action is determined for each grid compartment and host cell based on the local antibiotic concentrations (Figure 6C). PK and PD parameters were extensively calibrated to experimental in vitro and in vivo data (rabbits and non-human primates) on INH and RIF[75]. Our systems pharmacology approach allows us to probe antibiotic treatment in the context of a granuloma in ways not previously possible, integrating antibiotic activities, immune response dynamics, and spatio-temporal aspects of an evolving granuloma. Furthermore, we can explore host variation in both immune responses and PK, to provide a view of the host factors that contribute to the heterogeneous nature of TB and treatment. The model allows for the following unique analyses: true side-by-side comparison of different treatment regimens and dose sizes in the same granulomas; prediction of time to sterilization; identification of early indicators of treatment outcome; identification of key host mechanisms and antibiotic attributes controlling treatment outcome (in terms of percentage of granulomas sterilized, time to sterilization and final bacterial load in nonsterilized granulomas). Therefore this adapted PK/PD granuloma model is an excellent tool for suggesting possible improvements or alterations to current antibiotic treatments, as well as exploring a large number of dosing regimens and antibiotic combinations to narrow the search space for animal studies and clinical trials. We illustrate model outcomes for a representative granuloma treated with daily oral dosing of INH and RIF in Figure 7. We are able to evaluate granuloma PK, including average antibiotic concentrations for the granuloma as well as at specific locations in the granuloma. Average antibiotic concentrations remain below effective concentrations for the majority of dosing intervals inside granulomas (Figure 7A). Antibiotic concentration gradients form within granulomas, with lower concentrations, and therefore lower cumulative exposure, toward the center (Figure 7C and 7D). Simultaneous exposure to effective concentrations of both antibiotics inside the granuloma is infrequent. We note

 

9  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Integrative Biology

Page 10 of 37 View Article Online

411   412   413   414   415   416   417   418   419   420   421   422   423   424   425   426   427   428   429   430   431   432   433   434   435   436   437   438   439   440   441   442   443   444   445   446   447   448   449   450   451   452   453   454   455   456  

that monotherapy – exposure to effective concentrations of only a single antibiotic – can contribute to the development of or selection for drug resistant mutants[30, 78]. Model results also provide insight into the spatial and temporal bacterial response to treatment. The bacterial populations of TB disease are heterogeneous and complex. We represent this heterogeneity by modeling three different bacterial subpopulations: intracellular replicating, extracellular replicating and extracellular non-replicating bacteria. These subpopulations have different susceptibilities to INH and RIF[79, 80] (reflected in differing C50 (Figure 7A), Emax and Hill constant values), and we can track the response of each subpopulation to treatment in the representative granuloma (Figure 7B). Suboptimal antibiotic concentrations lead to bacterial growth between doses, likely a major factor contributing to the long treatment periods required for treating TB. As treatment progresses the intracellular and non-replicating extracellular bacterial subpopulations persist, while replicating extracellular populations are eliminated. In addition to a daily dosing regimen, the Centers for Disease Control and Prevention (CDC) also approve alternative dosing regimens of two or three times weekly 25. Analysis of 500 simulated granulomas predicts that this intermittent dosing increases both the time to sterilization (clearance) and the percentages of granulomas not sterilized for INH and RIF treatment alone or in combination[75]. This is contradictory to findings obtained recently using a model based on non-specific antibiotic parameters for treatment of self-limiting infections[81]. However, treatment outcomes are clearly both antibioticand pathogen-specific. In our model, pre-treatment infection severity (including bacterial burden, host cell activation and host cell death) and antibiotic exposure are predictive of treatment outcome. Our results suggest that both host and bacterial attributes continue to play important roles during antibiotic treatment. Finally, we note that our results are based on individual granuloma simulations, although the expectation is that granulomas that fail to clear bacteria could lead to active TB. 3. Inhaled Antibiotics As described above, current oral antibiotic regimens of RIF and INH lead to poor antibiotic penetration into granulomas causing sub-optimal exposure. This necessitates lengthy treatment durations causing chronic toxicity and concerns with patient compliance[2, 34]. A proposed alternative strategy to oral delivery of antibiotics for TB is inhaled delivery. In this delivery mode, fabricated carriers loaded with antibiotics are dosed into the lungs via an aerosol delivery system[82, 83]. Delivery of antibiotics via an inhaled route may overcome many limitations of oral dosing for treatment of TB by providing direct dosing to the infection site, reduced systemic toxicity and clearance, and improved patient compliance with reduced dosing frequency[82-85]. To rationally design inhaled formulations of antibiotics for TB treatment, it is necessary to understand the contributions of PK, PD, and behavior of the carriers (e.g. drug release) at the site of infection. Measuring and understanding these dynamics in clinically relevant models (e.g. non-human primates) is difficult and costly. Thus, systems pharmacology approaches are needed to quickly assess the efficacy and dynamics of inhaled formulations for the treatment of TB.

 

10  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Page 11 of 37

Integrative Biology View Article Online

457   458   459   460   461   462   463   464   465   466   467   468   469   470   471   472   473   474   475   476   477   478   479   480   481   482   483   484   485   486   487   488   489   490   491   492   493   494   495   496   497   498   499   500   501   502  

We extend our existing computational model of granuloma function and antibiotic treatment discussed above to include inhaled dosing and antibiotic release from a generalized carrier system [75] (Cilfone et al., submitted for publication, 2014). We modify the PK model to allow for dosing via both inhaled and oral routes by adding a non-infected lung compartment and an intracellular macrophage sub-compartment at pseudo-steady state (Figure 6B). Carriers are modeled as agents and behavior includes carrier movement, macrophage phagocytosis of carriers, dispersal from macrophages, and extra- and intracellular degradation (Figure 6A). In the non-infected lung and intracellular macrophage compartments, a homogenous representation of inhaled carriers is used. Release of antibiotics from carriers occurs in both the intra- and extracellular environment and is described by a diffusion-degradation equation with time varying boundary conditions[86-88]. We utilize the PD model constructed and calibrated in[75]. Using this model, we can begin to rationally design inhaled formulations of RIF and INH to be given at reduced dose frequencies (every two-weeks) with equivalent or better sterilizing capabilities as compared to conventional daily oral regimens. We can rapidly compare oral and inhaled doses, allowing us to assess whether existing antibiotics would be a promising candidates for inhaled formulations. We illustrate model predictions showing behavior of an inhaled dose of either RIF or INH given once every two weeks in Figure 8. Based on model sensitivity analysis, we identified possible inhaled formulations of INH and RIF that lead to equivalent or reduced bacterial loads at 7 days post-treatment initiation compared to daily oral formulations. For an inhaled formulation the total two-week dose (inhaled – 1x dose; oral – 14x doses) of INH is 12-fold lower compared to the oral formulation, while the total two-week dose for an inhaled formulation of RIF is the same as in the oral formulation. The model predicts that antibiotic concentrations in granulomas remain more stable over an entire dosing window (2-weeks) with inhaled formulations (Figure 8A) than with daily oral doses (Figure 7A). In the case of INH, average granuloma concentrations are sustained above C50 values for intra- and extracellular M. tuberculosis populations for the entire dosing window (Figure 8A). However, in the case of RIF, the inhaled formulation only eclipses the C50 of extracellular M. tuberculosis immediately after dosing and never surpasses the C50 for intracellular or non-replicating M. tuberculosis (Figure 8A). The average granuloma concentration of RIF slowly decreases, indicating that inhaled formulations cannot maintain effective concentrations of RIF over the two-week dosing window. We can also examine predicted treatment efficacy at the individual granuloma level for both drugs, comparing inhaled and oral formulations given once every two weeks and daily, respectively. There is no significant difference in successfully treated granulomas between the inhaled and oral formulations of RIF (Figure 8B). However, the inhaled formulation of INH sterilizes granulomas earlier than the oral formulation (Figure 8B). Treatment efficacy of inhaled formulations of RIF and INH, in comparison with their daily oral counterparts, is controlled by antibiotic concentrations mentioned above and the cumulative exposure in granulomas in a dosing window. A single inhaled dose of INH, given every two-weeks, leads to increased cumulative exposure in the granuloma

 

11  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Integrative Biology

Page 12 of 37 View Article Online

503   504   505   506   507   508   509   510   511   512   513   514   515   516   517   518   519   520   521   522   523   524   525   526   527   528   529   530   531   532   533   534   535   536   537   538   539   540   541   542   543   544   545   546   547   548  

compared to daily oral dosing (Figure 8C). A single inhaled dose of RIF, given every two-weeks, and daily oral dosing of RIF lead to similar cumulative exposure (Figure 8C). An inhaled formulation of RIF may not be practical because effective concentrations of RIF cannot be maintained for an entire dosing window, there are early increases in peripheral toxicity (defined as cumulative exposure in the peripheral compartment), and the required two-week total dose would have to exceed ~90% w/w in a polymeric carrier formulation (Cilfone et al., submitted for publication, 2014). However, RIF is one of many rifamycin antibiotics with differing PD [89]and PK properties. To further illustrate the potential capabilities of our approach, we tested how PD properties of RIF and other rifamycin antibiotics could be altered to improve the feasibility of an inhaled formulation. For instance, if a RIF derivative could be synthesized with different C50 characteristics would the efficacy of an inhaled formulation change? Using the same inhaled formulation of RIF as in Figure 8A-C, we varied the C50 values for the three modeled bacterial populations. As C50 values increase, the mean time to sterilize a granuloma remains constant or increases slightly (Figure 8D). When C50 values are decreased by ~ 50%, the mean time to sterilization decreases dramatically from ~80-100 days of treatment to ~3050 days of treatment (Figure 8D). Therefore, a RIF-derivative with different PD properties could make an inhaled formulation a practical possibility. We summarize the findings from our computational models with regard to oral and inhaled delivery of the first-line antibiotics INH and RIF in Figure 9. When designing treatment regimens and delivery systems, it is important to consider all relevant dynamics. INH and RIF distribute differently within the host and are eliminated at different rates, leading to very different dynamics at the host level and the site of infection (granuloma). For inhaled antibiotic delivery, host-level PK together with the dynamics of the delivery system (slow or fast release from the carrier) lead to their different dynamics at the site of infection. These influences must be considered together with the granuloma dynamics in designing therapeutics. 4. Vaccines The holy grail in TB therapy is the development of an effective vaccine (Figure 2 – intervention 6). When antigen is delivered to the body, antigen-presenting cells (APCs) (e.g. dendritic cells) present it in the context of MHC molecules to T cells circulating through LNs. T cells with specificity for that antigen/MHC complex bind, differentiate and proliferate to produce effector and memory T cells (Figure 1, right side). Central memory cells recirculate from blood to lymphoid organs, and can persist for years, awaiting activation by a second antigen challenge. Effector memory cells migrate to sites of infection. These cells have a shorter lifespan than central memory cells, but they can perform effector functions immediately after encountering a second antigen challenge. When vaccines are effective, these memory cells are able to protect an individual from disease. Perhaps not surprisingly, given the complexity of the host-pathogen dynamics, we do not yet understand what characteristics of an immune response correlate with protection against M. tuberculosis. Considering the high cost and time required to perform animal testing and human trials, computational models developed using a

 

12  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Page 13 of 37

Integrative Biology View Article Online

549   550   551   552   553   554   555   556   557   558   559   560   561   562   563   564   565   566   567   568   569   570   571   572   573   574   575   576   577   578   579   580   581   582   583   584   585   586   587   588   589   590   591   592   593   594  

systems biology approach can be an important supplement for hypothesis generation to aid TB vaccine design, especially in the early stages. To study how memory cells generated from vaccination could influence the course of M. tuberculosis infection, we incorporated two additional physiological compartments – LNs and blood – into our computational model of the site of infection (lung granuloma). This 3-compartmental physiological model tracks relevant cells and molecules that participate in generation of adaptive immunity and ensuing responses during M. tuberculosis infection (Figure 10). Building on our previous work[90], we use an ODE model to capture the dynamics of cells within LNs, although we note that to address questions requiring an understanding of spatial dynamics within LNs we have also developed agent-based models[91-93]. ODEs describe the evolution of naïve, precursor, central memory, effector memory, and effector T cells for both CD4+ and CD8+ T cells. Naïve and central memory cells can be recruited to LNs and are primed or activated at a rate based on the number of antigen-bearing DCs in the LN[94]. Shown in Figure 1 is the lymphatic system, whereby DCs traffic from sites of infection (here, lung) to LNs. To simplify, we assume that antigen-bearing DCs are recruited into LNs at a rate proportional to the number of macrophages that interacted with M. tuberculosis at that time step. We refer to this as the “APC proxy” (Figure 10). After priming, T cells enter a precursor pool, where they begin to proliferate. Cells in this state are not allowed to exit the LN due to the early activation markers they express[95]. Precursor cells eventually differentiate into central or effector T cells, and a portion of the effector T cells become effector memory cells. We also model a blood compartment, allowing immune cells to traffic from the LN to the site of infection where they can participate in the immune response. Blood is a well-mixed compartment, and therefore we use ODEs to represent the dynamics. The lung, LN, and blood compartment models are linked via cell trafficking terms, and physiological scaling is used to correctly account for the appropriate volumes of the compartments. An effective vaccine must trigger the immune response to generate a sufficient number of effector memory and central memory T cells that can act quickly in a recall response, preventing or controlling infection. Although the numbers required for successful protection are not known, our computational model can be used to generate predictions. Others have begun to explore this question as well, using mathematical modeling and bioinformatics approaches (www.epivax.com)[96]. For a simple illustration here, we assume that a vaccine will generate a particular level of M. tuberculosis-specific effector memory and central memory cells. These cells are assumed to be circulating postvaccination in the blood compartment. A recall response (via introduction of M. tuberculosis, as in our earlier models) is then simulated to test whether infection with M. tuberculosis is cleared, controlled, or neither. In other words, simulations may suggest what levels of memory cells are required for vaccine efficacy. We can compare “unvaccinated” with “vaccinated” cases to learn more about the protection that memory cells can provide. For the unvaccinated case, infected macrophages, T cells and bacteria progress into a contained granuloma with a relatively stable structure over time (Figure 11A, upper row), as seen with our earlier models (e.g.

 

13  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Integrative Biology

Page 14 of 37 View Article Online

595   596   597   598   599   600   601   602   603   604   605   606   607   608   609   610   611   612   613   614   615   616   617   618   619   620   621   622   623   624   625   626   627   628   629   630   631   632   633   634   635   636   637   638   639   640  

Figure 4B). If there are a sufficient number of memory cells present as a result of vaccination prior to the infection, the granuloma may not form, or may resolve quickly after a short period of growth (Figure 11A, lower row). To test how levels of different types of memory cells affect protection, we varied the initial condition for the numbers of effector memory and central memory classes of both CD4+ and CD8+ T cells that are present in the blood compartment. We introduce M. tuberculosis infection into the lung during a scenario where the parameters are biased toward a host phenotype that can form a granuloma that can contain infection, and track how the presence of circulating memory cells affects granuloma outcomes. Each setup is replicated 50 times, and the probability that a granuloma clears its bacterial load is counted (Figure 11B). We see that increasing memory CD4+ T cells does not influence the outcome of a granuloma. However, the chance of sterilization (clearance) increases when more memory CD8+ T cells are present, especially when a high proportion of them are effector memory cells. Even with this simple model, we see that an appropriate vaccine for M. tuberculosis could greatly alter the outcome of infection. The goal now is to design a vaccine that generates the necessary levels and ratios of memory cells. In a recent study, using an agent-based LN model, we predicted that the relative abundance of different T cell subsets could be tuned by controlling the quantity and quality of APCs[91]. These computational studies, together with bioinformatics analyses, animal vaccine, and human trials, are necessary to both improve our understanding of what is needed to develop a successful vaccine for TB and to help narrow the design space of possibilities. Discussion Although TB has been around for thousands of years, much is not understood about this complex infection. TB is a leading cause of death from infectious disease worldwide, second only to HIV-1/AIDS[1]. With a long and complicated antibiotic regimen required for TB treatment, there are a myriad of issues that can lead to treatment failure, including non-compliance, individual variations in antibiotic PK/PD, development of drug resistance, and differences among bacterial phenotypes. In recent years, several groups have taken a systems biology approach to identify critical metabolic and genetic regulatory pathways in the bacterium, with hopes of identifying new drug targets (e.g. [97-100] while others have focused on the alveolar macrophage host[101, 102]. These efforts have advanced our understanding of single cell level interactions and dynamics for both immune cells and pathogens. Our approach here complements that work but focuses on the multi-scale and multi-organ influences that determine infection outcomes in vivo. The advantage of this approach is that it allows us to understand the impact of a molecular-scale perturbation (e.g. in a rate or molecular concentration) at a granuloma, a tissue-scale readout; similarly the impact of events in the lymph node (e.g. memory cell generation) at a granuloma can be examined. These insights can help us understand how to narrow the design space for therapeutics including vaccines. We can simultaneously incorporate systems pharmacology approaches to describe how well a particular drug will reach and influence the target (e.g. a signaling pathway in a macrophage).

 

14  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Page 15 of 37

Integrative Biology View Article Online

641   642   643   644   645   646   647   648   649   650   651   652   653   654   655   656   657   658   659   660   661   662   663   664   665  

An in vivo systems biology approach to TB has much to offer to hypothesis generation and therapeutic design. Computational models can integrate information from experimental work focused on molecular, cellular, and tissue scales. Iteration between experiments and modeling is essential to building reliable computational models and designing appropriate experiments. New biological findings can easily be added to the computational model. For example, we are now including additional cell populations (T cell subsets and neutrophils) that new data suggest are important to granuloma dynamics. In addition to the antibiotic studies described herein, other combinations of first-line and second-line antibiotics can be studied using our drug model platform to allow for rapid screening of a wide and unwieldy drug regimen space. We are also exploring immunomodulation in tandem with antibiotics. This multiple-hit approach targeting immune responses while simultaneously limiting pathogen growth has great potential for success and could lead to patient-specific treatments, a goal of individualized medicine. Although we have focused here on lung granulomas and the processes that affect them, more work is needed to understand how infection status correlates with the numbers and characteristics of granulomas that are observed in vivo; forthcoming NHP and human data will be useful for this. Finally, there is an urgent need to identify biomarkers of infection status and progression in TB. This is particularly true in developing countries where resources are limited and the need to parse whom to treat, and when, is urgent. We currently are exploring using machine learning as biomarker discovery tool for TB. In vivo systems biology can play a major role in these important aspects of TB intervention.

 

15  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Integrative Biology

Page 16 of 37 View Article Online

666   667   668   669   670   671   672   673   674   675   676   677   678   679  

Acknowledgements. We thank Dr. Simone Marino and Dr. JoAnne Flynn for helpful discussions and insights, and Paul Wolberg and Joe Waliga for computational support. This research was supported by the following computational resources: the Open Science Grid (OSG), which is in supported by the National Science Foundation and the U.S. Department of Energy's Office of Science; the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by National Science Foundation grant number ACI-1053575; the National Energy Scientific Computing Center (NERSC), which is supported by the Office of Science of the U.S. Department of Energy under contract no. DE-AC02-05CH11231; the resources and services provided by Advanced Research Computing at the University of Michigan, Ann Arbor. This research was funded by NIH grants R01 EB012579 (DEK and JJL) and R01 HL 110811 (DEK and JJL). The authors declare no conflicts of interest.

 

16  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Page 17 of 37

Integrative Biology View Article Online

680   681   682   683   684   685   686   687   688   689   690   691   692   693   694   695   696   697   698   699   700   701   702   703   704   705   706   707   708   709   710   711   712   713   714   715   716   717   718   719   720   721   722   723   724  

Figure legends. Figure 1. Overview of the immune response to M. tuberculosis infection. M. tuberculosis replicates within macrophages. Some bacteria are killed via non-pathogen specific processes (innate immunity). Dendritic cells present antigen to naïve T cells in the lymph node, generating effector T cells (CD4+ and CD8+) that travel back to the site of infection to kill bacteria (adaptive immune response). Granulomas form in lungs as a result of these events. In non-human primates, granulomas range in size from ~1-6 mm in diameter (median value 2 mm)[10, 12]. Multiple granulomas are present in a single host and likely each one is seeded by a single bacterium[10, 103]. Memory T cells (CD4+ and CD8+) are also generated by processes in the lymph node. Figure 2. Multi-scale and multi-compartment view of host-pathogen dynamics during M. tuberculosis infection. Six potential interventions are also shown. Figure 3. Three elements of our computational approach to granuloma formation and function. (A) An agent-based model describes cellular actions. (B) Receptor binding, trafficking and signaling models are described with ODEs. (C) Molecular diffusion is described by partial differential equations. These model elements are linked, allowing information to be continually exchanged across scales. Figure 4. Granuloma model calibration and snapshots. (A) Comparison of CFU/lesion data from non-human primates (NHP) with computational model predictions (median – solid black line, min/max – dashed black lines). More detail on the model is given in [61]. NHP data from 32 animals collected between 28 and 296 days post-infection has been previously published in[10, 17]. (B) Sample simulation snapshot shows a granuloma that is containing infection at 200 days post-infection. (C) Sample simulation snapshot with different parameter values than in (B) shows a granuloma that fails to contain infection. Snapshot legend colors: resting macrophages (green), infected macrophages (orange), chronically infected macrophage (red), activated macrophage (dark blue), pro-inflammatory T cell (pink), cytotoxic T cell (purple), regulatory T cell (aqua), extracellular bacteria (brown), and caseation (cross-hatch). Figure 5. Simulated immunomodulation of IL-10 and TNF in granulomas. (A) Left side: Mean CFU per lesion for WT and IL-10 deletion (IL-10 K/O) lesions at 200 days postinfection. Percent of lesions becoming sterile by 200 days is indicated. Right side: Mean CFU per lesion for WT and IL-10 deletion (IL-10 K/O) lesions that were non-sterile at 200 days post-infection. Error bars indicate SD. (B) CFU for WT and IL-10 deletions starting at day 25, 50, 75, or 100 days post-infection. Percent of lesions becoming sterile by 200 days is indicated. Error bars indicate SD. (C) CFU per lesion and (D) Number of caseated compartments per lesion for granulomas with differing ratios of mean TNF to IL-10 concentrations. The ratio of TNF to IL-10 was modulated by increasing/decreasing rates of TNF and/or IL-10 production from all cell types. Individual circles represent individual lesions.

 

17  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Integrative Biology

Page 18 of 37 View Article Online

725   726   727   728   729   730   731   732   733   734   735   736   737   738   739   740   741   742   743   744   745   746   747   748   749   750   751   752   753   754   755   756   757   758   759   760   761   762   763   764   765   766   767   768   769   770  

Figure 6. Additions to the computational model of granuloma formation and function that allow for antibiotic dosing. (A) Granuloma PK of antibiotics described in the agentbased model include consideration of vascular permeability, diffusion and uptake by host cells. Inhaled antibiotics further include delivery particle deposition, movement and antibiotic release. (B) Host PK models describe movement of drug through the body and into the lung lesion (granuloma) using ODEs. The model for oral delivery (blue area) is expanded to allow inhaled delivery (blue + green areas). (C) PD are modeled independently of drug delivery method and are location- and bacterial subpopulationspecific. Killing rates are calculated using a Hill curve defined by the slope (Hill constant), maximum killing rate (Emax) and concentration where 50% activity is achieved (C50). These parameters are specific for different bacterial subpopulations. [75]. Figure 7. Simulated antibiotic treatment of a representative granuloma. The granuloma is allowed to form for 100 days, after which treatment is initiated with daily doses of INH (15 mg/kg) and RIF (20 mg/kg) for an additional 180 days. These doses give plasma PK similar to that seen in humans[104]. (A) Average INH and RIF concentrations in the granuloma shown in panel C and the corresponding total bacterial load in the granuloma over time. C50 values (see Figure 6) for INH and RIF are indicated by dashed lines for each bacterial subpopulation (blue – INH; red – RIF). (B) Bacterial subpopulations in the granuloma over time during 180 days of treatment. (C) Snapshot of the granuloma on day 100, before treatment starts. (D) Cumulative exposure of the granuloma in panel C to RIF (top left) and INH (bottom right) over the first week of treatment, showing spatial distribution. Color bars are scaled between 0 and the EC80 (exposure where 80% of max efficiency is achieved) for each antibiotic. Figure 8. Simulated antibiotic treatment of granulomas using inhaled formulations. Granulomas are allowed to form for 100 days, after which 200 days of treatment is simulated with inhaled formulations of INH and RIF dosed every two-weeks. (A) Mean INH (blue) and RIF (red) concentrations in the granuloma for the first 14-day dosing window. C50 values (see Figure 6) for INH and RIF are indicated by dashed lines for each bacterial subpopulation (blue – INH, red – RIF). (B) Percent of granulomas sterilized at indicated times after the initiation of treatment – INH (blue) and RIF (red). Inhaled formulations (solid lines) dosed every two-weeks are compared to daily oral dosing strategies (dashed lines). Granulomas still present at 300 days post-infection are considered failed treatments. (C) Mean INH (blue) and RIF (red) cumulative exposure in the granuloma for the first 14-day dosing window. Inhaled formulations dosed every twoweeks are compared to daily oral dosing strategies. (D) Mean time to granuloma sterilization when the C50 values (intracellular, extracellular, and non-replicating populations) of RIF are increased/decreased. RIF C50 values are given as a percentage of the original value. (A-C) * p ≤ 0.05, ** p ≤ 0.01, *** p ≤ 0.001, **** p ≤ 0.0001. INH – Inhaled (N = 81), Oral (N = 87). RIF – Inhaled (N = 83), Oral (N = 87). Figure 9. Antibiotic dynamics within granulomas are simultaneously influenced by host PK, granuloma PK, dosing regimens, and delivery route. Relative rates for INH and RIF are shown above (INH) or below (RIF) arrows. The transit compartment represents absorption in the gut and transit to systemic circulation. Oral and inhaled dosing regimens

 

18  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Page 19 of 37

Integrative Biology View Article Online

771   772   773   774   775   776   777   778   779   780   781   782   783   784   785   786   787   788   789   790   791   792   793   794   795   796   797   798   799   800   801   802   803   804   805   806   807   808   809   810  

and inhaled carrier release kinetics need to be designed with host PK and granuloma PK in mind. For INH, slow distribution to other organs, slow clearance, and low permeability allow for slow inhaled carrier release kinetics (all relative to RIF). For RIF, rapid distribution to other organs, rapid clearance, and high permeability must be compensated for by fast inhaled carrier release kinetics. Figure 10 Three-compartment model framework for simulating the influence of memory cells (which can be generated by vaccines) on granuloma formation and function. The lung (site of infection) is represented with our agent-based model (GranSim) as described in previous sections, and two ODE models capture LN and blood dynamics. T cells that are tracked in LNs include: CD4+ and CD8+ T cells, and each of these can be further classified into: N (Naïve), CM (Central Memory), EM (Effector Memory), P(precursor cells), E (Effector). APCs such as DCs circulate from the lung to the LN to prime the adaptive immune response. The CM, EM, E and N classes can travel between LN and blood compartments as indicated by arrows. Only E and EM (total effector class) can travel to the infection site in the lung. Both M. tuberculosis-specific and non-specific T cells are accounted for in our model. For our in-silico experiments, we changed the initial conditions of equations describing the number of different memory cells in the blood to represent the memory cells that we assume have been generated after vaccination (shown in box). The cell and bacterial time courses and granuloma spatial outcomes in the lung are tracked to assess the level of protection derived from the simulated vaccine. Figure 11 Simulated effects of immune memory on granuloma outcomes. (A) Snapshots of granuloma progression over time with or without memory cells generated from vaccination. When no memory cells are present at the beginning (top row), the site with an initial infected macrophage develops into a granuloma and maintains the structure through the 200 days of simulation. With sufficient memory cells circulating (bottom row), a granuloma appears briefly but quickly resolves as the infection ends in bacterial clearance (sterilization). (B) Infection is simulated with different combinations of initial conditions for each type of memory cell (central, effector, CD4+ and CD8+ T cells). Four groups of simulations are run, each with a fixed low (20 µL-1) or high (100 µL-1) concentration for total CD4+ or CD8+ T memory cells. Within each group, the ratio of central memory (CM) to effector memory (EM) are set to 9:1, 1:1, or 1:9. Each scenario is simulated 50 times, and the outcomes of each granuloma are assessed. Shown in the color are the proportions of simulations that ended in a granuloma that cleared all bacteria.

 

19  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Integrative Biology

Page 20 of 37 View Article Online

811   812   813   814   815   816   817   818   819   820   821   822   823   824   825   826   827   828   829   830   831   832   833   834   835   836   837   838   839   840   841   842   843   844   845   846   847   848   849   850   851   852   853   854  

References

1.   2.   3.   4.   5.   6.   7.   8.   9.   10.   11.   12.   13.   14.  

15.  

 

WHO,  Global  tuberculosis  report,  2014.   Dartois,  V.,  The  path  of  anti-­‐tuberculosis  drugs:  from  blood  to  lesions  to   mycobacterial  cells.  Nat  Rev  Microbiol,  2014.  12(3):  p.  159-­‐67.   Ramakrishnan,  L.,  Revisiting  the  role  of  the  granuloma  in  tuberculosis.   Nature  reviews.  Immunology,  2012.  12(5):  p.  352-­‐66.   Flynn,  J.L.  and  J.  Chan,  Immunology  of  tuberculosis.  Annu  Rev  Immunol,   2001.  19:  p.  93-­‐129.   O'Garra,  A.,  et  al.,  The  immune  response  in  tuberculosis.  Annual  review  of   immunology,  2013.  31:  p.  475-­‐527.   Flynn,  J.L.,  Lessons  from  experimental  Mycobacterium  tuberculosis   infections.  Microbes  Infect,  2006.  8(4):  p.  1179-­‐88.   Flynn,  J.L.  and  E.  Klein,  Pulmonary  Tuberculosis  in  Monkeys,  in  A  Color   Atlas  of  Comparative  Pulmonary  Tuberculosis  Histopathology,  V.D.  J.   Leong,  T.  Dick,  Editor.  2011,  CRC  Press,  Taylor  &  Francis.  p.  83-­‐106.   Lin,  P.L.  and  J.L.  Flynn,  Understanding  latent  tuberculosis:  a  moving   target.  J  Immunol,  2010.  185(1):  p.  15-­‐22.   Modlin,  R.L.  and  B.R.  Bloom,  TB  or  not  TB:  that  is  no  longer  the  question.   Sci  Transl  Med,  2013.  5(213):  p.  213sr6.   Lin,  P.L.,  et  al.,  Sterilization  of  granulomas  is  common  in  active  and  latent   tuberculosis  despite  within-­‐host  variability  in  bacterial  killing.  Nat  Med,   2014.  20(1):  p.  75-­‐9.   Lin,  P.L.,  et  al.,  Early  events  in  Mycobacterium  tuberculosis  infection  in   cynomolgus  macaques.  Infect  Immun,  2006.  74(7):  p.  3790-­‐803.   Lin,  P.L.,  et  al.,  Quantitative  comparison  of  active  and  latent  tuberculosis   in  the  cynomolgus  macaque  model.  Infect  Immun,  2009.  77(10):  p.  4631-­‐ 42.   Gideon,  H.P.  and  J.L.  Flynn,  Latent  tuberculosis:  what  the  host  "sees"?   Immunol  Res,  2011.  50(2-­‐3):  p.  202-­‐12.   Cilfone,  N.A.,  et  al.,  Multi-­‐scale  modeling  predicts  a  balance  of  tumor   necrosis  factor-­‐alpha  and  interleukin-­‐10  controls  the  granuloma   environment  during  Mycobacterium  tuberculosis  infection.  PloS  one,   2013.  8(7):  p.  e68680.   Marino,  S.,  et  al.,  TNF  and  IL-­‐10  are  major  factors  in  modulation  of  the   phagocytic  cell  environment  in  lung  and  lymph  node  in  tuberculosis:  a   next-­‐generation  two-­‐compartmental  model.  J  Theor  Biol,  2010.  265(4):   p.  586-­‐98.  

20  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Page 21 of 37

Integrative Biology View Article Online

855   856   857   858   859   860   861   862   863   864   865   866   867   868   869   870   871   872   873   874   875   876   877   878   879   880   881   882   883   884   885   886   887   888   889   890   891   892   893   894   895   896   897   898   899   900  

16.   17.   18.   19.   20.   21.   22.  

23.   24.   25.  

26.   27.   28.   29.   30.  

 

Wigginton,  J.E.  and  D.  Kirschner,  A  model  to  predict  cell-­‐mediated   immune  regulatory  mechanisms  during  human  infection  with   Mycobacterium  tuberculosis.  J  Immunol,  2001.  166(3):  p.  1951-­‐67.   Lin,  P.L.,  et  al.,  Radiologic  responses  in  cynomolgous  macaques  for   assessing  tuberculosis  chemotherapy  regimens.  Antimicrob  Agents   Chemother,  2013.   Guzzetta,  G.  and  D.  Kirschner,  The  roles  of  immune  memory  and  aging  in   protective  immunity  and  endogenous  reactivation  of  tuberculosis.  PloS   one,  2013.  8(4):  p.  e60425.   Barry,  C.E.,  3rd,  et  al.,  The  spectrum  of  latent  tuberculosis:  rethinking  the   biology  and  intervention  strategies.  Nat  Rev  Microbiol,  2009.  7(12):  p.   845-­‐55.   Fallahi-­‐Sichani,  M.,  et  al.,  Differential  risk  of  tuberculosis  reactivation   among  anti-­‐TNF  therapies  is  due  to  drug  binding  kinetics  and   permeability.  J  Immunol,  2012.  188(7):  p.  3169-­‐78.   Marino,  S.,  et  al.,  Differences  in  reactivation  of  tuberculosis  induced  from   anti-­‐TNF  treatments  are  based  on  bioavailability  in  granulomatous   tissue.  PLoS  Comput  Biol,  2007.  3(10):  p.  1909-­‐24.   Chakravarty,  S.D.,  et  al.,  Tumor  necrosis  factor  blockade  in  chronic   murine  tuberculosis  enhances  granulomatous  inflammation  and   disorganizes  granulomas  in  the  lungs.  Infect  Immun,  2008.  76(3):  p.   916-­‐26.   Harris,  J.,  J.C.  Hope,  and  J.  Keane,  Tumor  necrosis  factor  blockers   influence  macrophage  responses  to  Mycobacterium  tuberculosis.  The   Journal  of  infectious  diseases,  2008.  198(12):  p.  1842-­‐50.   Wallis,  R.S.,  et  al.,  Granulomatous  infectious  diseases  associated  with   tumor  necrosis  factor  antagonists.  Clin  Infect  Dis,  2004.  38(9):  p.  1261-­‐5.   Blumberg,  H.M.,  et  al.,  American  Thoracic  Society/Centers  for  Disease   Control  and  Prevention/Infectious  Diseases  Society  of  America:  treatment   of  tuberculosis.  American  journal  of  respiratory  and  critical  care   medicine,  2003.  167(4):  p.  603-­‐62.   Zhang,  Y.,  W.W.  Yew,  and  M.R.  Barer,  Targeting  persisters  for   tuberculosis  control.  Antimicrob  Agents  Chemother,  2012.  56(5):  p.   2223-­‐30.   Gomez,  J.E.  and  J.D.  McKinney,  M.  tuberculosis  persistence,  latency,  and   drug  tolerance.  Tuberculosis  (Edinb),  2004.  84(1-­‐2):  p.  29-­‐44.   Ahmad,  Z.,  et  al.,  The  potent  bactericidal  activity  of  streptomycin  in  the   guinea  pig  model  of  tuberculosis  ceases  due  to  the  presence  of  persisters.   J  Antimicrob  Chemother,  2010.  65(10):  p.  2172-­‐5.   Connolly,  L.E.,  P.H.  Edelstein,  and  L.  Ramakrishnan,  Why  is  long-­‐term   therapy  required  to  cure  tuberculosis?  PLoS  Med,  2007.  4(3):  p.  e120.   Mitchison,  D.A.,  How  drug  resistance  emerges  as  a  result  of  poor   compliance  during  short  course  chemotherapy  for  tuberculosis.  The   international  journal  of  tuberculosis  and  lung  disease  :  the  official   journal  of  the  International  Union  against  Tuberculosis  and  Lung   Disease,  1998.  2(1):  p.  10-­‐5.   21  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Integrative Biology

Page 22 of 37 View Article Online

901   902   903   904   905   906   907   908   909   910   911   912   913   914   915   916   917   918   919   920   921   922   923   924   925   926   927   928   929   930   931   932   933   934   935   936   937   938   939   940   941   942   943   944   945   946  

31.   32.   33.   34.   35.  

36.   37.   38.   39.   40.   41.   42.  

43.   44.   45.   46.  

47.  

 

Weis,  S.E.,  et  al.,  The  effect  of  directly  observed  therapy  on  the  rates  of   drug  resistance  and  relapse  in  tuberculosis.  The  New  England  journal  of   medicine,  1994.  330(17):  p.  1179-­‐84.   Chang,  K.C.,  et  al.,  Treatment  of  tuberculosis  and  optimal  dosing   schedules.  Thorax,  2011.  66(11):  p.  997-­‐1007.   Saltini,  C.,  Schedule  or  dosage?  The  need  to  perfect  intermittent  regimens   for  tuberculosis.  Am  J  Respir  Crit  Care  Med,  2006.  174(10):  p.  1067-­‐8.   Zumla,  A.,  P.  Nahid,  and  S.T.  Cole,  Advances  in  the  development  of  new   tuberculosis  drugs  and  treatment  regimens.  Nat  Rev  Drug  Discov,  2013.   12(5):  p.  388-­‐404.   Lin,  P.L.,  et  al.,  The  multistage  vaccine  H56  boosts  the  effects  of  BCG  to   protect  cynomolgus  macaques  against  active  tuberculosis  and   reactivation  of  latent  Mycobacterium  tuberculosis  infection.  The  Journal   of  clinical  investigation,  2012.  122(1):  p.  303-­‐14.   Weiner,  J.,  3rd  and  S.H.  Kaufmann,  Recent  advances  towards  tuberculosis   control:  vaccines  and  biomarkers.  Journal  of  internal  medicine,  2014.   275(5):  p.  467-­‐80.   Montagnani,  C.,  et  al.,  Vaccine  against  tuberculosis:  what's  new?  BMC   infectious  diseases,  2014.  14  Suppl  1:  p.  S2.   Kaufmann,  S.H.,  Tuberculosis  vaccines:  time  to  think  about  the  next   generation.  Seminars  in  immunology,  2013.  25(2):  p.  172-­‐81.   Aagaard,  C.,  et  al.,  A  multistage  tuberculosis  vaccine  that  confers  efficient   protection  before  and  after  exposure.  Nature  medicine,  2011.  17(2):  p.   189-­‐94.   Andersen,  P.  and  J.S.  Woodworth,  Tuberculosis  vaccines  -­‐  rethinking  the   current  paradigm.  Trends  in  immunology,  2014.   Churchyard,  G.J.,  et  al.,  Advances  in  immunotherapy  for  tuberculosis   treatment.  Clin  Chest  Med,  2009.  30(4):  p.  769-­‐82,  ix.   Rook,  G.A.,  D.B.  Lowrie,  and  R.  Hernandez-­‐Pando,  Immunotherapeutics   for  tuberculosis  in  experimental  animals:  is  there  a  common  pathway   activated  by  effective  protocols?  The  Journal  of  infectious  diseases,  2007.   196(2):  p.  191-­‐8.   Wallis,  R.S.,  Reconsidering  adjuvant  immunotherapy  for  tuberculosis.   Clin  Infect  Dis,  2005.  41(2):  p.  201-­‐8.   Fallahi-­‐Sichani,  M.,  et  al.,  Identification  of  key  processes  that  control   tumor  necrosis  factor  availability  in  a  tuberculosis  granuloma.  PLoS   Comput  Biol,  2010.  6(5):  p.  e1000778.   Tay,  S.,  et  al.,  Single-­‐cell  NF-­‐kappaB  dynamics  reveal  digital  activation   and  analogue  information  processing.  Nature,  2010.  466(7303):  p.  267-­‐ 71.   Braun,  D.A.,  M.  Fribourg,  and  S.C.  Sealfon,  Cytokine  response  is   determined  by  duration  of  receptor  and  signal  transducers  and   activators  of  transcription  3  (STAT3)  activation.  The  Journal  of   biological  chemistry,  2013.  288(5):  p.  2986-­‐93.   Giacomini,  E.,  et  al.,  Infection  of  human  macrophages  and  dendritic  cells   with  Mycobacterium  tuberculosis  induces  a  differential  cytokine  gene   22  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Page 23 of 37

Integrative Biology View Article Online

947   948   949   950   951   952   953   954   955   956   957   958   959   960   961   962   963   964   965   966   967   968   969   970   971   972   973   974   975   976   977   978   979   980   981   982   983   984   985   986   987   988   989   990   991  

48.   49.   50.   51.  

52.  

53.   54.   55.   56.   57.   58.   59.  

60.   61.  

 

expression  that  modulates  T  cell  response.  J  Immunol,  2001.  166(12):  p.   7033-­‐41.   Blomgran,  R.,  et  al.,  Mycobacterium  tuberculosis  inhibits  neutrophil   apoptosis,  leading  to  delayed  activation  of  naive  CD4  T  cells.  Cell  host  &   microbe,  2012.  11(1):  p.  81-­‐90.   Watson,  V.E.,  et  al.,  Apoptosis  in  mycobacterium  tuberculosis  infection  in   mice  exhibiting  varied  immunopathology.  The  Journal  of  pathology,   2000.  190(2):  p.  211-­‐20.   Qin,  H.,  et  al.,  SOCS3  deficiency  promotes  M1  macrophage  polarization   and  inflammation.  Journal  of  immunology,  2012.  189(7):  p.  3439-­‐48.   Cavnar,  S.P.,  et  al.,  Microfluidic  source-­‐sink  model  reveals  effects  of   biophysically  distinct  CXCL12  isoforms  in  breast  cancer  chemotaxis.   Integrative  biology  :  quantitative  biosciences  from  nano  to  macro,  2014.   6(5):  p.  564-­‐76.   Mattila,  J.T.,  et  al.,  Microenvironments  in  tuberculous  granulomas  are   delineated  by  distinct  populations  of  macrophage  subsets  and  expression   of  nitric  oxide  synthase  and  arginase  isoforms.  Journal  of  immunology,   2013.  191(2):  p.  773-­‐84.   Linderman,  J.J.  and  D.E.  Kirschner,  In  silico  models  of  M.  tuberculosis   infection  provide  a  route  to  new  therapies.  Drug  Discovery  Today:   Disease  Models,  2014(In  press).   Davis,  J.M.  and  L.  Ramakrishnan,  The  role  of  the  granuloma  in  expansion   and  dissemination  of  early  tuberculous  infection.  Cell,  2009.  136(1):  p.   37-­‐49.   Via,  L.E.,  et  al.,  Tuberculous  granulomas  are  hypoxic  in  guinea  pigs,   rabbits,  and  nonhuman  primates.  Infect  Immun,  2008.  76(6):  p.  2333-­‐ 40.   Fallahi-­‐Sichani,  M.,  et  al.,  A  systems  biology  approach  for  understanding   granuloma  formation  and  function  in  tuberculosis.,  in  Systems  biology  of   tuberculosis,  D.B.a.A.K.  J.  McFadden,  Editor.  2013,  Springer.   Fallahi-­‐Sichani,  M.,  et  al.,  Multiscale  computational  modeling  reveals  a   critical  role  for  TNF-­‐alpha  receptor  1  dynamics  in  tuberculosis   granuloma  formation.  J  Immunol,  2011.  186(6):  p.  3472-­‐83.   Fallahi-­‐Sichani,  M.,  D.E.  Kirschner,  and  J.J.  Linderman,  NF-­‐kappaB   Signaling  Dynamics  Play  a  Key  Role  in  Infection  Control  in  Tuberculosis.   Front  Physiol,  2012.  3:  p.  170.   Ray,  J.C.,  J.L.  Flynn,  and  D.E.  Kirschner,  Synergy  between  individual  TNF-­‐ dependent  functions  determines  granuloma  performance  for  controlling   Mycobacterium  tuberculosis  infection.  J  Immunol,  2009.  182(6):  p.  3706-­‐ 17.   Segovia-­‐Juarez,  J.L.,  S.  Ganguli,  and  D.  Kirschner,  Identifying  control   mechanisms  of  granuloma  formation  during  M.  tuberculosis  infection   using  an  agent-­‐based  model.  J  Theor  Biol,  2004.  231(3):  p.  357-­‐76.   Cilfone,  N.A.,  et  al.,  Computational  modeling  predicts  interleukin-­‐10   control  of  lesion  sterilization  by  balancing  early  host-­‐immunity-­‐mediated  

23  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Integrative Biology

Page 24 of 37 View Article Online

992   993   994   995   996   997   998   999   1000   1001   1002   1003   1004   1005   1006   1007   1008   1009   1010   1011   1012   1013   1014   1015   1016   1017   1018   1019   1020   1021   1022   1023   1024   1025   1026   1027   1028   1029   1030   1031   1032   1033   1034   1035  

62.   63.  

64.   65.   66.   67.   68.  

69.   70.   71.   72.  

73.   74.   75.  

 

antimicrobial  responses  with  caseation  during  mycobacterium   tuberculosis  infection.  J  Immunol,  2014.   Marino,  S.,  et  al.,  Macrophage  polarization  drives  granuloma  outcome   during  Mycobacterium  tuberculosis  infection.  Infect  Immun,  2014.   Cilfone,  N.A.,  D.E.  Kirschner,  and  J.J.  Linderman,  Strategies  for  efficient   numerical  implementation  of  hybrid  multi-­‐scale  agent-­‐based  models  to   describe  biological  systems  (Accepted  pending  minor  revisions).  Cellular   and  Molecular  Bioengineering,  2014.   Kirschner,  D.E.,  et  al.,  Tuneable  resolution  as  a  systems  biology  approach   for  multi-­‐scale,  multi-­‐compartment  computational  models.  Wiley   Interdiscip  Rev  Syst  Biol  Med,  2014.  6(4):  p.  289-­‐309.   Marino,  S.,  J.J.  Linderman,  and  D.E.  Kirschner,  A  multifaceted  approach   to  modeling  the  immune  response  in  tuberculosis.  Wiley  Interdiscip  Rev   Syst  Biol  Med,  2011.  3(4):  p.  479-­‐89.   Marino,  S.,  et  al.,  A  methodology  for  performing  global  uncertainty  and   sensitivity  analysis  in  systems  biology.  J  Theor  Biol,  2008.  254(1):  p.  178-­‐ 96.   Flynn,  J.L.,  et  al.,  An  essential  role  for  interferon  gamma  in  resistance  to   Mycobacterium  tuberculosis  infection.  J  Exp  Med,  1993.  178(6):  p.  2249-­‐ 54.   Lin,  P.L.,  et  al.,  Tumor  necrosis  factor  neutralization  results  in   disseminated  disease  in  acute  and  latent  Mycobacterium  tuberculosis   infection  with  normal  granuloma  structure  in  a  cynomolgus  macaque   model.  Arthritis  Rheum,  2010.  62(2):  p.  340-­‐50.   Ouyang,  W.,  et  al.,  Regulation  and  functions  of  the  IL-­‐10  family  of   cytokines  in  inflammation  and  disease.  Annu  Rev  Immunol,  2011.  29:  p.   71-­‐109.   Moore,  K.W.,  et  al.,  Interleukin-­‐10  and  the  interleukin-­‐10  receptor.  Annu   Rev  Immunol,  2001.  19:  p.  683-­‐765.   Bogdan,  C.,  Y.  Vodovotz,  and  C.  Nathan,  Macrophage  deactivation  by   interleukin  10.  J  Exp  Med,  1991.  174(6):  p.  1549-­‐55.   Balcewicz-­‐Sablinska,  M.K.,  H.  Gan,  and  H.G.  Remold,  Interleukin  10   produced  by  macrophages  inoculated  with  Mycobacterium  avium   attenuates  mycobacteria-­‐induced  apoptosis  by  reduction  of  TNF-­‐alpha   activity.  J  Infect  Dis,  1999.  180(4):  p.  1230-­‐7.   Higgins,  D.M.,  et  al.,  Lack  of  IL-­‐10  alters  inflammatory  and  immune   responses  during  pulmonary  Mycobacterium  tuberculosis  infection.   Tuberculosis  (Edinb),  2009.  89(2):  p.  149-­‐57.   Redford,  P.S.,  P.J.  Murray,  and  A.  O'Garra,  The  role  of  IL-­‐10  in  immune   regulation  during  M.  tuberculosis  infection.  Mucosal  Immunol,  2011.   4(3):  p.  261-­‐70.   Pienaar,  E.,  et  al.,  A  computational  tool  integrating  host  immunity  with   antibiotic  dynamics  to  study  tuberculosis  treatment.  Journal  of   Theoretical  Biology,  2014.  In  Press.  

24  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Page 25 of 37

Integrative Biology View Article Online

1036   1037   1038   1039   1040   1041   1042   1043   1044   1045   1046   1047   1048   1049   1050   1051   1052   1053   1054   1055   1056   1057   1058   1059   1060   1061   1062   1063   1064   1065   1066   1067   1068   1069   1070   1071   1072   1073   1074   1075   1076   1077   1078   1079   1080   1081  

76.   77.   78.   79.   80.   81.   82.   83.   84.   85.   86.   87.   88.   89.   90.   91.   92.  

 

Savic,  R.M.,  et  al.,  Implementation  of  a  transit  compartment  model  for   describing  drug  absorption  in  pharmacokinetic  studies.  J  Pharmacokinet   Pharmacodyn,  2007.  34(5):  p.  711-­‐26.   Regoes,  R.R.,  et  al.,  Pharmacodynamic  functions:  a  multiparameter   approach  to  the  design  of  antibiotic  treatment  regimens.  Antimicrob   Agents  Chemother,  2004.  48(10):  p.  3670-­‐6.   Loddenkemper,  R.,  D.  Sagebiel,  and  A.  Brendel,  Strategies  against   multidrug-­‐resistant  tuberculosis.  Eur  Respir  J  Suppl,  2002.  36:  p.  66s-­‐ 77s.   Jayaram,  R.,  et  al.,  Pharmacokinetics-­‐pharmacodynamics  of  rifampin  in   an  aerosol  infection  model  of  tuberculosis.  Antimicrob  Agents   Chemother,  2003.  47(7):  p.  2118-­‐24.   Jayaram,  R.,  et  al.,  Isoniazid  pharmacokinetics-­‐pharmacodynamics  in  an   aerosol  infection  model  of  tuberculosis.  Antimicrob  Agents  Chemother,   2004.  48(8):  p.  2951-­‐7.   Ankomah,  P.  and  B.R.  Levin,  Exploring  the  collaboration  between   antibiotics  and  the  immune  response  in  the  treatment  of  acute,  self-­‐ limiting  infections.  Proc  Natl  Acad  Sci  U  S  A,  2014.  111(23):  p.  8331-­‐8.   Sosnik,  A.,  et  al.,  New  old  challenges  in  tuberculosis:  potentially  effective   nanotechnologies  in  drug  delivery.  Adv  Drug  Deliv  Rev,  2010.  62(4-­‐5):  p.   547-­‐59.   Muttil,  P.,  C.  Wang,  and  A.J.  Hickey,  Inhaled  drug  delivery  for  tuberculosis   therapy.  Pharm  Res,  2009.  26(11):  p.  2401-­‐16.   Misra,  A.,  et  al.,  Inhaled  drug  therapy  for  treatment  of  tuberculosis.   Tuberculosis  (Edinb),  2011.  91(1):  p.  71-­‐81.   Griffiths,  G.,  et  al.,  Nanobead-­‐based  interventions  for  the  treatment  and   prevention  of  tuberculosis.  Nat  Rev  Microbiol,  2010.  8(11):  p.  827-­‐34.   Siepmann,  J.  and  F.  Siepmann,  Mathematical  modeling  of  drug  delivery.   Int  J  Pharm,  2008.  364(2):  p.  328-­‐43.   Kanjickal,  D.G.  and  S.T.  Lopina,  Modeling  of  drug  release  from  polymeric   delivery  systems-­‐-­‐a  review.  Crit  Rev  Ther  Drug  Carrier  Syst,  2004.  21(5):   p.  345-­‐86.   Arifin,  D.Y.,  L.Y.  Lee,  and  C.H.  Wang,  Mathematical  modeling  and   simulation  of  drug  release  from  microspheres:  Implications  to  drug   delivery  systems.  Adv  Drug  Deliv  Rev,  2006.  58(12-­‐13):  p.  1274-­‐325.   Burman,  W.J.,  K.  Gallicano,  and  C.  Peloquin,  Comparative   pharmacokinetics  and  pharmacodynamics  of  the  rifamycin   antibacterials.  Clin  Pharmacokinet,  2001.  40(5):  p.  327-­‐41.   Marino,  S.,  M.  El-­‐Kebir,  and  D.  Kirschner,  A  hybrid  multi-­‐compartment   model  of  granuloma  formation  and  T  cell  priming  in  Tuberculosis.  J   Theor  Biol,  2011.  280(1):  p.  50-­‐62.   Gong,  C.,  Linderman,  JJ  and  Kirschner,  DE.,  Harnessing  the  heterogeneity   of  T  cell  differentiation  fate  to  fine-­‐tune  generation  of  effector  and   memory  T  cells.  Frontiers  in  Immunology,  2013.  (in  press).   Gong,  C.,  et  al.,  Predicting  lymph  node  output  efficiency  using  systems   biology.  Journal  of  theoretical  biology,  2013.  335C:  p.  169-­‐184.   25  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Integrative Biology

Page 26 of 37 View Article Online

1082   1083   1084   1085   1086   1087   1088   1089   1090   1091   1092   1093   1094   1095   1096   1097   1098   1099   1100   1101   1102   1103   1104   1105   1106   1107   1108   1109   1110   1111   1112   1113   1114   1115   1116   1117   1118  

93.  

Linderman,  J.J.,  et  al.,  Characterizing  the  dynamics  of  CD4+  T  cell  priming   within  a  lymph  node.  J  Immunol,  2010.  184(6):  p.  2873-­‐85.   94.   Girard,  J.P.  and  T.A.  Springer,  High  endothelial  venules  (HEVs):   specialized  endothelium  for  lymphocyte  migration.  Immunol  Today,   1995.  16(9):  p.  449-­‐57.   95.   Shiow,  L.R.,  et  al.,  CD69  acts  downstream  of  interferon-­‐alpha/beta  to   inhibit  S1P1  and  lymphocyte  egress  from  lymphoid  organs.  Nature,  2006.   440(7083):  p.  540-­‐4.   96.   Bhunu,  C.P.,  et  al.,  Modelling  the  effects  of  pre-­‐exposure  and  post-­‐ exposure  vaccines  in  tuberculosis  control.  J  Theor  Biol,  2008.  254(3):  p.   633-­‐49.   97.   Jamshidi,  N.  and  B.O.  Palsson,  Investigating  the  metabolic  capabilities  of   Mycobacterium  tuberculosis  H37Rv  using  the  in  silico  strain  iNJ661  and   proposing  alternative  drug  targets.  BMC  systems  biology,  2007.  1:  p.  26.   98.   Slayden,  R.A.,  et  al.,  Updating  and  curating  metabolic  pathways  of  TB.   Tuberculosis  (Edinb),  2013.  93(1):  p.  47-­‐59.   99.   Beste,  D.J.,  et  al.,  (1)(3)C  metabolic  flux  analysis  identifies  an  unusual   route  for  pyruvate  dissimilation  in  mycobacteria  which  requires   isocitrate  lyase  and  carbon  dioxide  fixation.  PLoS  pathogens,  2011.  7(7):   p.  e1002091.   100.   Galagan,  J.E.,  et  al.,  The  Mycobacterium  tuberculosis  regulatory  network   and  hypoxia.  Nature,  2013.  499(7457):  p.  178-­‐83.   101.   Rao,  K.S.,  D.  Kumar,  and  S.  Mande,  Probing  Gene  Regulatory  Networks  to   Decipher   Host–Pathogen  Interactions,  in  Systems  biology  of  tuberculosis,  D.B.a.A.K.  J.   McFadden,  Editor.  2013,  Springer.   102.   Beste,  D.J.,  et  al.,  (13)C-­‐Flux  Spectral  Analysis  of  Host-­‐Pathogen   Metabolism  Reveals  a  Mixed  Diet  for  Intracellular  Mycobacterium   tuberculosis.  Chemistry  &  biology,  2013.  20(8):  p.  1012-­‐21.   103.   Ragheb,  M.N.,  et  al.,  The  mutation  rate  of  mycobacterial  repetitive  unit   loci  in  strains  of  M.  tuberculosis  from  cynomolgus  macaque  infection.   BMC  Genomics,  2013.  14:  p.  145.   104.   Lin,  P.L.,  et  al.,  Metronidazole  prevents  reactivation  of  latent   Mycobacterium  tuberculosis  infection  in  macaques.  Proc  Natl  Acad  Sci  U   S  A,  2012.  109(35):  p.  14188-­‐93.    

 

26  

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

Page 27 of 37

Integrative Biology View Article Online

Figure 1.        

Trafficking to Lymph Node ~8-12 Days

Inhaled Bacterium M. tuberculosis

Lymph Node

Macrophage Dendritic Cell

Lymphatics and Bloodstream

Innate Immunity

Adaptive Immunity

T Cell Priming Memory T Cells (Effector and Central)

Lungs Granuloma Formation

   

 

 

Effector T Cells (Pro-Inflammatory, Cytotoxic, and Trafficking From Lymph Node ~14-17 Days Regulatory)

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

DOI: 10.1039/C4IB00295D

 

Integrative Biology

Page 28 of 37 View Article Online

DOI: 10.1039/C4IB00295D

Figure 2.

Lymph node

Lung

5

Lymphatics

Organ scale

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

 

Blood

Granuloma

Tissue scale Granuloma

6

5

Cellular scale

TNF Gradient

Mtb

2 1

TNF

3

Molecular scale

4 IL-10 NFκB Signaling

Symbols Memory T-cell

M. tuberculosis

Activated macrophage

Mtb antigen

Infected macrophage

TNF receptor

Resting macrophage

IL-10 receptor TNF

Dendritic cell

 

Interventions ( ) 1) NFκB signaling

5) Memory T cells (vaccine)

2) TNF receptor internalization

6) Antibiotics

3) TNF concentration 4) IL-10 concentration

IL-10

 

 

Page 29 of 37

Integrative Biology View Article Online

DOI: 10.1039/C4IB00295D

A. Agent Based Model Cellular Apoptosis

+

Macrophage Bursting

Macrophage Activation Infection

Chemotactic-Based Movement

Uptake of Mtb

Chronic Infection

Cellular Recruitment From Vasculature Bacterial Growth

TNF

∂C ∆2 =D C ∂t

v1

v2 v4

v6

v3 v12

v9 v8

IL-10

v5

v13

v11 v10

v7

B. Receptor Trafficking and Signaling Model

C. Soluble Molecule Diffusion Model

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

Figure 3.

CFU / Lesion

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

A

107 B

106

100

NHP Model

105

104

103

102

101

0

Days (post-infection) 100 200 300 1mm 1mm

Integrative Biology Accepted Manuscript

Integrative Biology View Article Online

Page 30 of 37

DOI: 10.1039/C4IB00295D

Figure 4.

C

Page 31 of 37

Integrative Biology View Article Online

DOI: 10.1039/C4IB00295D

Figure 5. B 56%

ns

***

500

CFU / Lesion (200 Days Post-Infection)

CFU / Lesion (200 Days Post-Infection)

29%

Non-Sterile Lesions

400 300 200 100 0

WT

IL-10 K/O

WT

IL-10 K/O

600

Percentage of Sterile Lesions 12%

48%

84%

54%

400

*** ****

200

0

30%

****

WT

0

25

50

75

D 104

60

Caseation (# / Lesion)

103 102

40

20

101 Sterile 0.1

1

[TNF] / [IL-10]

0

**

****

Start of IL-10 Depletion (Days Post-Infection)

C

24%

0.1

1

[TNF] / [IL-10]

100

Integrative Biology Accepted Manuscript

%of Sterile Lesions

CFU / Lesion (200 Days Post-Infection)

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

A

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

Integrative Biology View Article Online

Page 32 of 37

DOI: 10.1039/C4IB00295D

Figure 6.

Page 33 of 37

Integrative Biology View Article Online

DOI: 10.1039/C4IB00295D

1 0.1

C50,BE C50,BI

0.01

400 200

0.001 0.0001

600

102

105

108

150

225

B

Simulated granuloma at day 100 post infection

150

200

250

300

Time (Days)

Time (Days)

C

200

0

0 300

Intracellular Replicating Extracellular Non-replicating Extracellular Total

400

D

Cumulative exposure to RIF and INH from day 100 to 107 post infection RIF

RIF

INH

INH

Integrative Biology Accepted Manuscript

C50,BI 1000 C50,BN C50,BE 800 C50,BN

CFU per granuloma

INH and RIF concentration in granuloma (mg/L)

A

CFU per granuloma

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

Figure 7.

Integrative Biology

Page 34 of 37 View Article Online

DOI: 10.1039/C4IB00295D

Figure 8.

C50,BE C50,BN

10-1 C50,BE C50,BI

10-2 100

107

114

p = 0.0022

75 50 25 0

0

80

RIF ns

****

40 0

Inhaled Oral

Inhaled Oral

100

150

200

250

300

D 250

Time to Sterilization (Days Post-Infection)

550 450 350 250

INH

INH - Inhaled INH - Oral RIF - Inhaled RIF - Oral

Time (Days Post-Infection)

Time (Days Post-Infection)

C

p = ns

200 150 100 0

Before Antibiotic Treatment 0

50

100

150

C50 Values (% of RIF Value)

200

Integrative Biology Accepted Manuscript

100

100

Failed Treatments

C50,BI C50,BN

Pre-Treatment

101

Percent of Successfully Treated Granulomas

INH and RIF Granuloma Concentration (mg/L)

B

Cumulative 14-Day Antibiotic Exposure (mg hr/L)

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

A

Page 35 of 37

Integrative Biology View Article Online

DOI: 10.1039/C4IB00295D

Transit

Low Permeability

Slow Fast

Slow

Body

Plasma

High Permeability

First-Pass and Non-Absorption

Lung

Fast

Granuloma

Clearance and Metabolism

[INH]

Oral Dose

Time

Slow Release

Fast Release

Isoniazid

Inhaled Oral

Rifampin

Inhaled Oral

Inhaled Dose

Time

Integrative Biology Accepted Manuscript

Antibiotic Pharmacokinetics

[RIF]

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

Figure 9.

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

Integrative Biology View Article Online

Page 36 of 37

DOI: 10.1039/C4IB00295D

Figure 10.

Integrative Biology Accepted Manuscript

Published on 20 April 2015. Downloaded by University of California - San Francisco on 28/04/2015 00:50:42.

Page 37 of 37 Integrative Biology DOI: 10.1039/C4IB00295D

View Article Online

Figure 11.

A multi-scale approach to designing therapeutics for tuberculosis.

Approximately one third of the world's population is infected with Mycobacterium tuberculosis. Limited information about how the immune system fights ...
9MB Sizes 0 Downloads 10 Views